Download Second-order intermodulation mechanisms in CMOS downconverters

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Analog television wikipedia , lookup

Multimeter wikipedia , lookup

Test probe wikipedia , lookup

Josephson voltage standard wikipedia , lookup

TRIAC wikipedia , lookup

Oscilloscope history wikipedia , lookup

Integrating ADC wikipedia , lookup

Ohm's law wikipedia , lookup

Phase-locked loop wikipedia , lookup

Superheterodyne receiver wikipedia , lookup

Amplifier wikipedia , lookup

Index of electronics articles wikipedia , lookup

Wien bridge oscillator wikipedia , lookup

Surge protector wikipedia , lookup

Analog-to-digital converter wikipedia , lookup

Regenerative circuit wikipedia , lookup

Power MOSFET wikipedia , lookup

Voltage regulator wikipedia , lookup

Current source wikipedia , lookup

Transistor–transistor logic wikipedia , lookup

Schmitt trigger wikipedia , lookup

Wilson current mirror wikipedia , lookup

Valve audio amplifier technical specification wikipedia , lookup

Power electronics wikipedia , lookup

Operational amplifier wikipedia , lookup

CMOS wikipedia , lookup

Switched-mode power supply wikipedia , lookup

Resistive opto-isolator wikipedia , lookup

Radio transmitter design wikipedia , lookup

Current mirror wikipedia , lookup

Valve RF amplifier wikipedia , lookup

Heterodyne wikipedia , lookup

Rectiverter wikipedia , lookup

Opto-isolator wikipedia , lookup

Transcript
394
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 38, NO. 3, MARCH 2003
Second-Order Intermodulation Mechanisms
in CMOS Downconverters
Danilo Manstretta, Member, IEEE, Massimo Brandolini, Student Member, IEEE, and Francesco Svelto, Member, IEEE
Abstract—An in-depth analysis of the mechanisms responsible for second-order intermodulation distortion in CMOS
active downconverters is proposed in this paper. The achievable
second-order input intercept point (IIP2) has a fundamental
limit due to nonlinearity and mismatches in the switching stage
and improves with technology scaling. Second-order intermodulation products generated by the input transconductor or due
to self-mixing usually contribute to determine the IIP2 even
though they can, at least in principle, be eliminated. The parasitic
capacitance loading the switching-stage common source plays
a key role in the intermodulation mechanisms. Moreover, the
paper shows that, besides direct conversion and low intermediate
frequency (IF), even superheterodyne receivers can suffer from
second-order intermodulation if the IF is not carefully chosen.
The test vehicle to validate the proposed analysis is a highly linear
0.18- m direct-conversion CMOS mixer, embedded in a fully
integrated receiver, realized for Universal Mobile Telecommunications System applications.
Index Terms—Active mixers, CMOS analog integrated circuits,
dc offset, direct conversion, IIP2, low IF, mismatch, radio receivers,
second-order distortion, self-mixing, UMTS.
I. INTRODUCTION
D
IRECT conversion and low-intermediate-frequency (IF)
architectures have gained increasing attention for the realization of fully integrated wireless receivers. The main reason
is that they lead to potentially low-cost solutions. Depending
on the application and on the specific standard, either of the
two can be better suited. Direct conversion suffers from drawnoise [1], [2], while low-IF
backs such as dc offset and
does not. In fact, a low-IF solution downconverts the signal at
a near-zero-IF frequency, in such a way that the offset can be
noise has less impact on receiver seneasily removed and
sitivity. On the other hand, in a direct-conversion solution, the
image-rejection requirements are not demanding because the
image is the signal itself. Conversely, to keep the image-rejection requirement within affordable levels, the intermediate frequency of a low-IF architecture is usually chosen in such a way
that the image channel is the adjacent channel, in which the
maximum allowed signal power is kept much lower than in the
other channels. As a consequence, second-order intermodulation products, whose bandwidth is twice the bandwidth of the
interferer, can impair the sensitivity of a low-IF solution. When
Manuscript received July 25, 2002; revised October 25, 2002. This work was
supported by the Italian National Program Pope.
D. Manstretta is with Agere Systems, Berkeley Heights, NJ 07922 USA
(e-mail: [email protected]).
M. Brandolini and F. Svelto are with the Department of Electronics, University of Pavia, 27100 Pavia, Italy.
Digital Object Identifier 10.1109/JSSC.2002.808310
second-order intermodulation constitutes a major problem, such
as in frequency-division-duplexing systems, neither of the two
architectures is intrinsically advantageous. In such cases, the realization of a highly integrated low-cost solution might even
be prevented. Even more, a superheterodyne architecture can
suffer from second-order intermodulation, if the IF frequency
is not carefully chosen. In fact, as it will be shown, a modulated blocker at a frequency halfway between received-signal
and local-oscillator (LO) frequencies would translate at IF, becoming indistinguishable from the signal.
This paper addresses the problem of second-order intermodulation distortion in highly integrated CMOS receivers. Usually, the downconverter determines the achievable second-order
input intercept point (IIP2) of the receiver. The focus is, then,
on the physical mechanisms of second-order intermodulation in
CMOS active mixers. In fact, only some of the mechanisms are
well known and no quantitative analysis exists. Experimental
results from a 0.18- m CMOS mixer, embedded in a fully integrated receiver for the Universal Mobile Telecommunications
System (UMTS), are discussed. The paper is organized as follows. Section II is a review of second-order intermodulation
distortion mechanisms, Section III is focused on self-mixing,
and Sections IV and V discuss the intrinsic phenomena, i.e., active devices nonlinearity and mismatches. Section VI presents
the design and experimental results of the highly linear UMTS
downconverter, and Section VII presents the conclusions.
II. MECHANISMS OF SECOND-ORDER
INTERMODULATION DISTORTION
One mechanism responsible for second-order intermodulation distortion is known as self-mixing [3]. It is due to parasitic
coupling of the radio-frequency (RF) signal into the LO port
and to a non-hard-switching – characteristic of the commutating stage. As a result, the mixer behaves as a multiplier and a
signal proportional to the square of the input signal is produced
at mixer output, resulting in base-band content.
A second mechanism is due to second-order nonlinearity in
the active devices of the transconductor [3]–[5]. The low-frequency intermodulation current at the output of the transconductor leaks to the output of the mixer without frequency conversion due to a duty-cycle distortion of the LO. As described
in [5], this effect is to first-order cancelled using a double-balanced topology. However, a more important source of leakage
exists that is not cancelled in double-balanced mixers. This is
due to mismatches between the current switching stage devices.
The combined effect of nonlinearity in the transconductor and
mismatches in the current switching stage will be discussed in
Section IV.
0018-9200/03$17.00 © 2003 IEEE
MANSTRETTA et al.: SECOND-ORDER INTERMODULATION MECHANISMS IN CMOS DOWNCONVERTERS
Fig. 1. Single balanced mixer used to model self-mixing.
395
(a)
Finally, second-order intermodulation products are generated
by nonlinearity in the switching stage. In this case, intermodulation products generation and leakage are processes which are
intimately related, because leakage at mixer output is due to
mismatches in the same devices. At low frequencies, the main
nonlinear effect of the switching pair is a nonlinear partition of
the current between the two active devices during the switching
transients [6]. Nonetheless, the achievable IIP2 is still very high.
On the other hand, at high frequencies, due to the presence of a
parasitic capacitance at the common source, the linearity drops
dramatically. This mechanism determines the ultimate limit to
the intrinsic mixer IIP2 and will be described in Section V.
(b)
III. SELF-MIXING
Let us consider the single balanced mixer, shown in Fig. 1. In
the ideal case of hard-switching behavior, the – characteristic
of the pair is shown in Fig. 2(a) and, assuming a square-wave
voltage applied at the switching pair, the differential output curfor one half of the period and
rent is equal to
for the other half. In this case, even if the input voltage is
coupled to the switching-stage LO input, the drain current of the
input transistor MRF is steered by either device of the pair for
exactly 50% of the period and no second-order product arises at
mixer output. In actuality, the – characteristic is not stepwise
and the applied LO signal is not a square wave. To gain an intuitive insight into the self-mixing mechanism, we will assume
a soft-switching – characteristic [shown in Fig. 2(b)] and a
sinusoidal LO signal. The mixer differential output current
is given by
(1)
where
(2)
,
depends on the input device current only.
For
,
depends on the product of the input deFor
vice current times the actual voltage signal at the switching-pair
input. As a result, when both switching-pair devices are simultaneously on, the output current contains a term proportional to
the input signal square.
Fig. 2. I –V characteristic of the switching pair: (a) hard-switching and
(b) soft-switching models.
The second-order intermodulation product in the differential
can be expressed as
output current
(3)
If we assume a large amplitude sinusoidal LO, i.e.,
with
, the switching time in, i.e., the interval during which
, is
terval
. Hence,
is obtained
given by
times a square-wave
as the product of
toggling between 1 and 0 with a duty cycle equal
function
.
,
, and
are
to
plotted in Fig. 3.
Let us consider the case where a strong out-of-band modulated blocker is received together with the desired signal. To
gain insight in the analysis, the modulated blocker is represented
by a double sideband suppressed carrier (DSB-SC) signal, i.e.,
.
,
Recognizing that
,
can be evaluated as
for
(4)
396
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 38, NO. 3, MARCH 2003
Fig. 5. IIP2 due to self-mixing versus LO power, with 2 mA biasing current
and device width as a parameter. Theoretical (solid line) and simulated (dots)
results.
Fig. 3. Time-domain waveforms in a soft-switching single balanced mixer:
= I
I , and self-mixing
applied sinusoidal LO V , output current I
function sw(t).
0
that all the mechanisms described in this paper also affect superheterodyne architectures.
Recalling the definition of IIP2 for DSB-SC signals, i.e., the
input power for which the second-order intermodulation products and the linear terms are equal and the mixer transconduc, the following extance conversion gain is given by
pression results:
IIP2 (dBm)
Fig. 4. Effect of self-mixing in a superheterodyne architecture, assuming a
modulated blocker (represented by a DSB-SC signal) halfway between RF and
LO frequencies.
where spectral components at higher frequencies, not of interest
in this analysis, have been neglected.
Equation (4) tells that, as far as direct or low-IF conversions
are concerned, the second term is responsible for intermodulation, while the first one is a dc offset, and the other spectral
components can be easily filtered out. Moreover, the last three
terms tell us that even a superheterodyne architecture suffers
from second-order intermodulation. In fact, if a strong blocker is
at a frequency halfway between the received-signal and LO frequencies, both the received-signal and second-order intermodulation products are at IF at mixer output. This undesirable situation is illustrated in Fig. 4, where the output spectrum shows
the received signal downconverted at IF, together with all the
low-frequency second-order intermodulation products.
In the following, we concentrate on direct-conversion and
low-IF architectures, where the problem of second-order intermodulation cannot be circumvented. However, it can be shown
(dBm).
(5)
The IIP2 due to self-mixing depends only on LO amplitude
and coupling, while it is, at least to first-order, independent of
any design parameter. To maximize the IIP2, the RF-to-LO coupling has to be minimized, as expected, and the LO amplitude
has to be maximized. Equation (5) has been verified through
simulation of the single balanced mixer of Fig. 1. A DSB-SC
signal is injected into the RF input port, while a 40-dB attenuated differential replica, representing RF-to-LO coupling, is injected into the LO port. Fig. 5 reports IIP2 versus LO power, for
mA. Fig. 6 reports IIP2 versus LO
various widths and
m. In both
power, for various biasing currents and W
cases, a very good agreement between simulated and theoretical
values is evident for LO powers higher than about 2 dBm. For
lower LO powers, the mixer behaves more like a linear multiand the simulations
plier. The conversion gain is lower than
start to deviate from the calculated values.
IV. TRANSCONDUCTOR NONLINEARITY
SWITCHING-PAIR MISMATCHES
AND
Even if the RF and LO ports were perfectly de-coupled, a
mixer would present second-order intermodulation distortion.
Besides the desired RF signal, low-frequency second-order
intermodulation products appear in the transconductor output
current spectrum. After the current-switching stage, the
signal is downconverted. The low-frequency intermodulation
products are upconverted around the LO frequency. However, mismatches in the current switches cause part of the
low-frequency current leaking at mixer output still around the
baseband. The combination of second-order nonlinearity in the
MANSTRETTA et al.: SECOND-ORDER INTERMODULATION MECHANISMS IN CMOS DOWNCONVERTERS
397
where
and
are the differential and common-mode second-order intermodulation currents at transconductor output, respectively, and
is the rms value of the low-frequency leakage gain.
and
produce second-order intermoduBoth
is
lation in the differential output current. Moreover,
converted into a differential voltage signal by mismatches in
the mixer load. Therefore, the rms differential output voltage
is given by
(8)
Fig. 6. IIP2 due to self-mixing versus LO power, with 300 m device width
and pair biasing current as a parameter. Theoretical (solid line) and simulated
(dots) results.
where is the mixer-load differential resistance and
resistance standard deviation. To calculate
,
, and need to be determined.
is the
in (8),
A. Transconductor Nonlinearity
Let us consider the Volterra series expansion of
and
,
common-mode and differential currents (
respectively) versus differential input voltage of the transconductor
(9)
From (9), low-frequency second-order common-mode and
differential output currents are given by
and
, respectively. If the two devices
is zero. On the other
are perfectly matched,
hand, taking into account mismatches in the two devices, due
or
C
variations,
to threshold voltage
is a random process and its standard deviation
is given by
Fig. 7. Double balanced mixer schematic.
transconductor and mismatches in the current switches gives
rise to second-order intermodulation products at mixer output.
This mechanism will be discussed referring to the double balanced mixer, reported in Fig. 7. The inherent second-order nonlinear terms in the MOS – characteristic give rise to secondorder intermodulation components in the current of each branch
) at transconductor output. Assuming low frequency
(
and
for the switching pairs M3–M4 and M5–M6,
leakage
respectively, the following intermodulation in the mixer output
results:
current
(6)
Considering that the two switching pairs have the same nominal design values but mismatches and, as a consequence,
are random processes, statistically independent, the root-mean) of
is given by
square (rms) value (
(7)
(10)
and
are the
where
and ,
second-order transconductance sensitivities to
is the standard deviation of
and
is the
respectively.
standard deviation of .
The derivation of the second-order nonlinear transconductances for three different implementations of the input stage
(fully differential, pseudodifferential, and LC degenerated
is given
pairs) is reported in Appendix A. The admittance
(taking into account the finite output
by
,
impedance of the current source), , and
respectively. A summary of the results is reported in Table I,
where is the second-order transconductance of the single de. The following interesting conclusions
vice and
can be drawn.
1) Given the typical matching parameters of a deep-submicron technology, threshold voltage mismatches determine
.
398
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 38, NO. 3, MARCH 2003
G
, @G
TABLE I
=dV , AND @G
=d FOR FULLY DIFFERENTIAL,
PSEUDODIFFERENTIAL, AND LC-DEGENERATED INPUT TRANSCONDUCTORS
Fig. 9. Second-order nonlinearity in pseudodifferential and LC degenerated
transconductors. Theoretical (solid line) and simulated (dots) G =G
versus
biasing current, for various device widths.
the common-mode second-order transconductance
,
as a function of the biasing current and for different device
widths, for a pseudodifferential transconductor. Both simulated
results and theoretical analysis results are reported together,
showing a very good agreement. The linearity improves with
increasing overdrive voltages, but it is significantly smaller
than in the fully differential topology case (by about 30 dB).
As a result, a fully differential topology is best suited to
achieve higher second-order linearity, ultimately limited by
mismatches in the threshold voltage of the devices.
Fig. 8. Second-order nonlinearity in a fully differential transconductor.
versus biasing current,
Theoretical (solid line) and simulated (dots) G =G
for various device widths.
2) In a fully differential solution, the second-order differential transconductance is typically higher than the
mV
common-mode one. In fact, assuming
value,
. On the other
and 1-mV
hand, with a typical 4-mA biasing current and 5-k
current-source output resistance,
.
3) In pseudodifferential and LC-degenerated topologies, the
common-mode second-order transconductance is much
is the
larger than the differential one. Moreover,
same for the two topologies.
4) The total second-order transconductance of a pseudodifferential topology is larger than that of a fully differential
one.
The results of the analysis have been verified through
simulations. The MOS model proposed in [7] has been used
for calculations. Fig. 8 plots the linear transconductance over
as
differential second-order transconductance ratio
a function of the biasing current and for different device widths,
for a fully differential transconductor. The two sets of curves
refer to mismatches in the beta coefficient and in the threshold
voltage at 3 . In both cases, the agreement between theory
and simulation is very good. The results in Fig. 8 confirm that
threshold voltage mismatch actually determines second-order
differential nonlinearity in a fully differential transconductor.
Linearity improves with increasing biasing currents and decreasing device widths, i.e., for larger device overdrive voltages.
Fig. 9 plots the ratio between the linear transconductance and
B. Switching-Pair Low-Frequency Leakage Gain
In a mixer with perfectly matched switching-pair devices,
no second-order intermodulation product is transferred at the
is upconverted while
remains
output because
a common-mode signal. Threshold voltage and variations determine a mismatch in the two devices, giving rise to a low-frequency leakage gain. The circuit shown in Fig. 10(a), showing
one single switching pair of the double balanced mixer, is used
to quantify , the leakage gain rms value. The device mismatch
is taken into account by means of an equivalent offset voltage
, in series with the switching-stage LO input.
source
Two distinct mechanisms have been identified. The first is
due to duty-cycle distortion in the output current waveform, and
the second is induced by the parasitic capacitance loading the
switching pair. The former will be referred to as direct (with
the direct leakage gain), the latter as indirect (with
the indirect leakage gain), by analogy with the mechanisms of
noise in a mixer [8].
Neglecting the parasitic capacitance in the single balanced
mixer of Fig. 10(a), the output current, normalized to the
input one, with an applied sinusoidal LO signal, is shown in
Fig. 10(b). The offset voltage is responsible for duty-cycle
distortion and the output current has a nonzero average value.
The output current waveform can be looked at as the input
current modulated by a square-wave toggling between 1 with
50% duty cycle plus a train of pulses with height equal to 2,
.
is the average value of the
width , and period
train of pulses
(11)
MANSTRETTA et al.: SECOND-ORDER INTERMODULATION MECHANISMS IN CMOS DOWNCONVERTERS
Fig. 11.
LO.
(a)
399
Equivalent model of the mixer in Fig. 10(a), assuming a square-wave
where
is the time constant when the biasing current is .
to the time constant is a time-varying
The sensitivity of
function, while the sensitivity of the time constant to the biasing current is assumed to be constant. To gain insight in this
intermodulation mechanism, the analysis is carried on in the frequency domain. The source voltage is found by inspection of the
circuit. Its derivative with respect to is given by
(14)
(b)
Fig. 10. (a) Single balanced mixer with offset voltage modeled in series with
the gate. (b) Waveforms versus time: applied LO (V ), actual output current
decomposed in output current without offset plus a train of pulses due to offset.
The offset voltage is a random variable, and its standard deviais given by
tion
From (13) and (14), the current signal produces two sidein the source voltage spectrum. Corbands around
respondingly, two sidebands are generated in the capacitor current spectrum and finally downconverted at a low frequency .
This means that a low-frequency second-order intermodulation
product at transconductor output is transmitted at mixer output
at the same frequency, even if the applied LO is a square-wave
signal. The resulting mixer second-order intermodulation lowis given by
frequency output current at
(12)
and
are the variance of and , respectively.
where
Using a large LO amplitude reduces the direct leakage gain
and as a consequence second-order intermodulation products
at mixer output. Ultimately, a square-wave input LO with in. Nonetheless,
finite slope at zero crossing makes
second-order products at mixer output are still present due to
another low-frequency gain mechanism. When a square-wave
LO signal is applied at the switching stage, the two devices are
alternately on during each half period. The effect of the offset
voltage can be analyzed exactly in the same way as the effect
of the equivalent noise of the switching-pair devices [8]. This
means that the pair can be looked at as a single source follower,
with an applied gate-voltage signal toggling between 0 and
(Fig. 11). The tail capacitor charges and discharges exponen, between 0 and
.
tially, with a time constant
Let us consider the effect of the low-frequency intermodulation
coming from the transconductor, represented by a small current
. Because the current
modulates the
signal
device biasing current, the charging-discharging time constant
will also be modulated. The source voltage can be expressed as
(13)
(15)
has been introduced.
where the relation
,
As expected, assuming no parasitic capacitance, i.e.,
. Actually, in a direct-conversion mixer,
noise
considerations call for large-area devices, biased at low currents,
leading to a relatively long time constant . Even assuming a
square-wave LO (i.e., no direct leakage), second-order intermodulation is produced and the indirect mechanism determines
the achievable IIP2. Before deriving quantitative conclusions,
we still have to complete the analysis, and consider the case of
an applied sine-wave LO. The theory developed so far is still
valid, though we need to consider that the source voltage varies
in response to an applied large sinusoid. In order to simplify the
nonlinear analysis, we furthermore assume the following.
1) Each device of the switching pair is on for half of the
period, i.e., when the corresponding voltage is positive.
2) The switching time is negligible with respect to the LO
period.
3) The operating point of the active transistors is constant
during each LO semiperiod.
400
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 38, NO. 3, MARCH 2003
Assuming, as in the former case, a low-frequency tone
, the output low-frequency second-order intermoduis found to be
lation current
(19)
Fig. 12. Equivalent model of the mixer in Fig. 10(a), taking into account the
effect of an applied sinusoidal LO.
This is reasonable, at least for
. However, we can expect some deviation from this very simple model. In fact, when
the LO peak amplitude is small, the switching time is not negligible with respect to the LO period. On the other hand, when
the LO peak amplitude is large, a high common-mode current is
injected into the common-source parasitic capacitance at twice
the LO frequency. This is likely to substantially modify the operating point of the active transistor during the LO semiperiod.
Nonetheless, this simple model is able to capture all the essential
mechanisms involved in the intermodulation process, giving acceptable quantitative accuracy. The key advantage with respect
to more complicated approaches is, then, a much easier understanding of the underlying mechanisms. The overall gate voltage
is given by the superposition of the rectified LO plus the periodic square-wave offset. The effect of the applied sinusoidal LO
will be studied referring to Fig. 12. The rectified sinusoid
can be expressed by means of its Fourier series expansion
(16)
is found by inspection of the circuit.
The source voltage
The following source voltage sensitivity to the time constant
) results:
(
(17)
Again, a low-frequency current signal, due to second-order
nonlinearity in the input transconductor, modulates the source
voltage via time constant modulation. The time constant has
content at dc and at all of the even harmonics of the LO frequency. The dc term has no effect because the corresponding
current in the capacitor is null. The terms at frequencies
are indeed important. In fact, referring to the mixer output versus
input current relation, reported in Fig. 10(b), we realize that the
train of pulses originating the direct leakage is also responsible
. The train of pulses
can
for a finite transmission at
be expressed as
(18)
From (15) and (19), the output second-order intermodulation
currents, due to source-voltage sidebands around even as well as
odd harmonics of the LO fundamental frequency, are both important. This is not surprising even though they are originated
by intermodulation of the low-frequency current tone with impressed signals, having magnitudes that are largely different. In
fact, the large LO sinusoid determines sidebands around
proportional to the peak amplitude
, while the switching-pair
. On the other hand, the sideleakage is proportional to
are directly proportional to the offset
bands around
and downconverted by the switching stage (i.e.,
voltage
with maximum conversion gain).
By means of (11), (15), and (19), the overall leakage gain
is found to be
(20)
From (20), the direct and the two indirect gains can add
, i.e., the overall
or subtract depending on the value of
leakage gain magnitude can be lower than the leakage gain
magnitude due to each single term. As a consequence, de, i.e., the biasing conditions at a given LO
pending on
frequency, the IIP2 can be higher than due to each separate
mechanism.
To validate the analysis, the circuit in Fig. 10(a) has been simulated, injecting a low-frequency tone, at the switching-stage
is 1.5 mA, the offset
common source. The biasing current
is 2 mV, the LO frequency is 2.1 GHz, and the simvoltage
ulation has been repeated for various LO oscillation amplitudes
and parasitic capacitance at the common source. Fig. 13 reports
the simulation results together with (20) versus the parasitic
capacitance. The maximum capacitance shown corresponds to
, i.e., a value where the assumptions leading to (20)
are still reasonable. Finally, notice that the series in (20) is trun. The agreement between simulations and calcucated for
lations is good, especially for moderate LO amplitudes, where
the model is expected to be more accurate. More remarkably,
Fig. 13 shows, in the entire considered range, a monotonic decrease of the leakage gain, due to canceling between the dif-
MANSTRETTA et al.: SECOND-ORDER INTERMODULATION MECHANISMS IN CMOS DOWNCONVERTERS
Fig. 13. Theoretical and simulated leakage gain versus parasitic capacitance
at common source, for various LO powers.
ferent mechanisms. In particular, for small capacitance, the direct leakage gain dominates. Increasing the capacitance value
determines a reduction of the leakage because the indirect mechanisms subtract. A reduction of 7 dB, with respect to the direct
leakage value, is predicted with 0.6-pF source capacitance for
6-dBm LO power.
Based on the above analysis, an estimate of the achievable
IIP2 due to transconductor nonlinearity and switching pair mismatches can be done. From (8), the input intercept second-order
) is given by
rms voltage point (
401
to the mixing operation, and determines a fundamental limit to
the achievable IIP2.
In Section IV, it has been shown that a low-frequency
current tone modulating the switching-stage transconductance
determines sidebands around
and all its multiples, in
the source voltage spectrum. The corresponding sidebands
in the capacitor current spectrum are downconverted at the
baseband by the switching stage. To describe the intermodulation distortion mechanisms associated with the switching
stage, the model assumed in Section IV-B (Fig. 12) can still
be used, provided higher order transconductances of active
devices, responsible for further intermodulation products generation, are taken into account. A DSB-SC RF current signal
injected into the switching stage
gives rise to source voltage spectral components at
and
, due to the transconductance modulation. More, , and
over, components at linear combinations of
are expected, due to higher order transconductance components
with
modulation. The most important are those at
, because they result in base-band content at
mixer output. The detailed circuit solution leading to the IIP2
analytical expression is reported in Appendix B. The output
second-order intermodulation currents due to side bands around
and
,
odd and even harmonics of the LO (
respectively) are given by
(21)
Even considering a mixer with a pseudodifferential transconductor, i.e., the worst-case topology for second-order nonlinearity performances, IIP2 values in excess of 65 dBm can be
m
m
achieved. In fact, let us consider, as an example,
transconductor devices, biased at about 5 mA. From (21),
dBV (i.e., IIP2
dBm) can be
Figs. 9 and 13, IIV2
obtained, provided the LO amplitude is in the order of 6 dBm,
and the mismatch between the load resistors is kept below
. The use of a fully differential transconductor potentially allows achieving much higher IIP2 values. In this case,
the output resistances matching requirement is more relaxed
because the second-order common-mode current is lower.
V. SWITCHING-PAIR NONLINEARITY AND MISMATCHES
Self-mixing and transconductor nonlinearity, together with
switching-pair mismatches, can limit the achievable IIP2 of a
mixer. Nonetheless, the two phenomena are not intrinsic to the
mixer operation, i.e., both can be eliminated, at least in principle. Self-mixing depends on coupling from RF to LO ports
and its effect can be eliminated via isolation techniques. AC
coupling between the transconductor and the switching stage
allows filtering out the transconductor low-frequency current,
which originates second-order intermodulation products.
Still another source of second-order intermodulation distortion exists. This is associated with the switching pair, is intrinsic
(22)
and
(23)
where and are the second-order and third-order transconductance, respectively.
, no
From (22) and (23), we conclude that for
second-order product is generated. This means that at low frequency, for small parasitic capacitance and large biasing currents, the switching stage gives rise to negligible second-order
intermodulation products at mixer output. This is not the case
when the transconductor is responsible for second-order intermodulation components generation. In fact, the low-frequency
second-order intermodulation current at transconductor output
is transmitted at the same frequency due to switching-pair duty-
402
Fig. 14. Theoretical and simulated IIP2 due to the switching stage. A typical
20-mS input transconductance is assumed. The series in equations (22) and (23)
are truncated for k = 2.
cycle distortion even for
. This leads to the interesting
conclusion that while the achievable IIP2 due to the transconductor will not improve with technology scaling, that due to the
switching stage will. Because the latter sets the actual intrinsic
limit to the mixer IIP2, technology scaling will allow larger
achievable IIP2.
To verify (22) and (23), the output second-order intermodulation current in the circuit of Fig. 10(a) has been simulated, injecting a DSB-SC current signal at the switching-stage source.
The simulation is repeated for various biasing currents. The corresponding mixer IIP2, assuming 20 mS input stage transconductance, is estimated. Fig. 14 shows both simulations and theoretical calculations versus biasing current. Notice that in the
theoretical plot of Fig. 14, the series in (22) and (23) are trun. The agreement is reasonably good in the entire
cated for
current range. The discrepancy is mainly attributed to the inaccuracy of the model, where, for example, the linear and higher
orders transconductances have been assumed constant during
the entire period. On the other hand, taking into account that
all the parameters are actually periodically time variant would
not allow arriving at an analytical solution in explicit form [6].
From Fig. 14, we observe that the IIP2 achieves a local maximum value at low currents, where the distinct mechanisms tend
to cancel. Notice that taking into account terms of order higher
in the series leads to a worse agreement with simulathan
tions. The reason is that higher order terms in the series are actually overestimated, because the assumption of instantaneous
switching between the pair devices is less and less valid at higher
frequencies. Finally, Fig. 15 shows the simulated IIP2 for different pair device widths. The effect is a slightly different current value where canceling between different mechanisms takes
place.
From the analysis carried on in the previous two sections,
we conclude that while second-order products generated by the
input transconductor can be made negligible, by using a fully
differential topology or by means of filtering, the intrinsic limit
to the achievable IIP2 is determined by the switching stage.
When a pseudodifferential topology is used for the transconductor, both the switching pair and the input stage concur to
determine the IIP2, and the achievable value depends on the details of the design.
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 38, NO. 3, MARCH 2003
Fig. 15. Simulated IIP2 due to the switching stage versus biasing current for
different device widths. The input stage transconductance is assumed to be
20 mS.
VI. MIXER DESIGN AND EXPERIMENTAL RESULTS
The test vehicle for the proposed analysis is a direct-conversion 0.18- m CMOS mixer designed for UMTS applications
[9]. Because UMTS is a frequency division duplexing system,
it requires a highly linear receiver to reject the transmitter
leakage, which constitutes a huge interfering signal 130 MHz
away from the receiver band. When direct-conversion architecture is adopted, both IIP2 and IIP3 requirements are very
stringent. Moreover, to meet the sensitivity requirement, UMTS
is also demanding in terms of noise figure. The mixer design
noise and high
is then particularly tough because low
linearity are usually contrasting requirements. In particular, the
noise in
switching-pair devices are an important source of
a direct-conversion CMOS mixer in the gigahertz range [8].
Their contribution to the noise figure is reduced using relatively
large areas and low biasing currents [10]. On the other hand,
low noise (i.e., large transconductance) and high linearity in the
input transconductor are simultaneously achieved only at the
expense of current consumption. To reflect these requirements,
the input-stage and switching-stage currents are set independently. Fig. 16 shows the diagram of the mixer chosen in this
design. The input transconductor is made of a pMOS fully differential input stage shunting an LC-degenerated nMOS stage.
In this way, the switching-stage biasing current is the difference
between the nMOS and the pMOS currents. To maximize the
input-stage IIP2, fully differential transconductors actually
offer the best performance. Nonetheless, we have chosen an
LC-degenerated nMOS stage for the following reasons.
1) Still in a 0.18- m CMOS technology, the switching-stage
nonlinearity constitutes the most important limitation to
the achievable IIP2.
2) The low supply voltage of 1.8 V is not sufficient to efficiently realize two shunt fully differential transconductors.
Moreover, an LC-degenerated stage has a higher IIP3 for
given biasing current.
To reduce third-order nonlinearity due to the switching
stage, an LC filter resonating at twice the LO frequency has
been introduced between the common source and ground. The
filter determines low common-mode impedance at twice the
LO frequency, i.e., reduces the common-mode oscillation at the
MANSTRETTA et al.: SECOND-ORDER INTERMODULATION MECHANISMS IN CMOS DOWNCONVERTERS
Fig. 16.
Fig. 17.
UMTS receiver block diagram.
Fig. 18.
Chip photograph.
403
Schematic of the mixer for UMTS.
switching-pair common source, increasing IIP3 [6]. Moreover,
it allows using larger LO amplitudes with a beneficial effect on
IIP2, due to the direct mechanism. Although the results reported
in Sections IV and V do not automatically apply to the mixer in
Fig. 16, it can be shown that the mechanisms originating IIP2
are exactly the same, the results are qualitatively the same, and
analytical expressions can be found in a quite straightforward
way. The impact of the filter on gain and noise performances
is limited because, for differential signals, it is equivalent to a
small capacitance.
As mentioned above, the proposed mixer has been embedded in a receiver front-end for UMTS applications. The
front-end block diagram, comprising a low-noise amplifier
(LNA), a quadrature mixer, and a variable-gain amplifier
(VGA), is shown in Fig. 17. A classical inductively degenerated
LC-loaded LNA provides 16-dB gain. After the mixers, VGAs
further amplify the signal by 19 dB. A low-frequency loop
around the VGA cancels any dc offset and creates a high-pass
transfer function with a cutoff frequency of 3 kHz. The chip was
fabricated in a 0.18- m six-metal-layers CMOS technology,
from STMicroelectronics. Fig. 18 shows the chip photograph.
Particular care was taken to have highly symmetric quadrature signal paths. It was also very important to have a cross
symmetry between the RF and LO lines so that differential
RF-to-LO and LO-to-RF couplings were minimized. The LO is
externally supplied.
For testing purposes, the available test pins are RF input,
mixer LO inputs (terminated externally with 50- differential
resistors), mixer outputs, and VGA outputs. Notice that this prototype is appropriate to prove the analysis of IIP2 in mixers, presented in this paper, because low-frequency second-order products generated by the LNA are completely rejected at mixer
input, due to the LNA resonant load.
All the measurements have been performed while the circuit
is supplied by a 1.8-V source. The RF-to-LO isolation is 65 dB.
Equation (5) suggests that the receiver IIP2 due to self-mixing is
around 70–75 dBm for LO powers ranging from 1 to 4 dBm.
This means that, given 16 dB of LNA gain, the IIP2 at mixer
input is in the range of 86–91 dBm. Of course, this is not the
only mechanism determining the mixer IIP2. To prove the validity of the theory proposed in Section III, a fraction of the RF
signal has been purposely injected into the LO input and the
output IM2 measured as a function of the LO power. The two
tones, representative of the transmitter leakage, are at 1980 and
1980.5 MHz, resulting in an output second-order intermodulation at 0.5 MHz. The received signal is at 2111 MHz and the LO
is at 2110 MHz, with an output signal at 1 MHz. Fig. 19 shows
measured and theoretical IIP2 versus LO power, with 28 dB
RF-to-LO signal ratio. For high LO powers, a very good agreement is evident. For low LO powers, a small difference is observed, as expected from circuit simulations.
Prior to the characterization of the mixer IIP2, due to the intrinsic mechanisms, the load resistances have been measured. At
the time of the design, not all the mechanisms determining the
IIP2 and described in this paper were completely understood.
In particular, the impact of the resistor accuracy on IIP2 was not
quantitatively clear. The measured relative accuracy of 1/600
is in good agreement with the design, but does not allow IIP2
values higher than about 66 dBm. Because higher peak values
are expected, perfect load match has been achieved by shunting
404
Fig. 19. Measured and theoretical IIP2 versus LO power with
RF-to-LO coupling.
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 38, NO. 3, MARCH 2003
028-dB
Fig. 21. Measured IIP2 versus switching-stage biasing current for various LO
powers.
To complete the receiver characterization, noise figure
and IIP3 measurements have been performed. The receiver
out-of-band IIP3, which is mainly determined by the mixer,
measured as specified by UMTS, is 2 dBm. Finally, the
receiver noise figure (integrated from 10 kHz to 1.92 MHz)
is 6.2 dB. The receiver front-end performances meet the
requirement specified by UMTS.
VII. CONCLUSION AND DISCUSSION
Fig. 20. Measured and simulated mixer IIP2 versus switching-stage biasing
current. The LO power is 6 dBm.
external resistors. The same frequency plan used for self-mixing
characterization has been used to characterize the intrinsic IIP2.
Fig. 20 shows the measured and simulated mixer IIP2 versus
the switching-pair biasing current with 5-dBm LO power. The
standing current in the pMOS transconductor is kept constant
and equal to 1.8 mA. Notice that the simulations at large currents
were performed using very low-load resistors. As expected, at
low current values, the switching pairs’ indirect mechanisms determine the IIP2, while at high currents, the IIP2 is limited by
the nonlinearity in the transconductor. The agreement between
measurements and simulations is very good in the entire range,
even though measured values peak in a slightly narrower current range. Values higher than 65 dBm are measured in the range
1.1–1.6 mA, with a maximum value as high as 75 dBm. Considering that the LNA gain is 16 dB, the mixer satisfies, in the
above current range, the IIP2 requirement of a UMTS system,
specified as about 48 dBm at antenna. Fig. 21 shows the measured IIP2 versus current for different LO powers. Values higher
than 60 dBm are measured in the range 1–1.7 mA.
The systematic approach in the analysis of second-order
products generations in mixers, proposed in this paper, has
disclosed interesting insights into the underlying mechanisms.
Quantitative results for self-mixing as well as for the intrinsic
mechanisms have been derived. In particular, the paramount
importance of the parasitic capacitance loading the switching
stage has been discussed in depth, for the first time. The paper
employs a simple equivalent circuit for the current-switching
stage, first proposed in [8], even though more accurate device
modeling is required to capture the intermodulation phenomena
responsible for second-order intermodulation products than is
required for noise analysis.
Both self-mixing and second-order products generated in the
input transconductor are not intrinsic to the mixer operation
and can be eliminated, at least in principle. On the other hand,
second-order products generated by the switching stage are
intrinsic and determine the maximum achievable IIP2. From
(22) and (23), the switching-stage nonlinearity determines the
second-order products at mixer output only in the presence of
parasitic capacitance loading the stage. This means not only
that at lower frequency this effect is less important, but also that
the intrinsic IIP2 is expected to increase at the given frequency
in more scaled technologies.
APPENDIX A
Let us consider the input-stage differential pair of the double
balanced mixer, reported in Fig. 7. The Taylor series expansion
of the – characteristic of device M is expressed as
(A.1)
where
and
are, respectively, the gate and source voltages.
MANSTRETTA et al.: SECOND-ORDER INTERMODULATION MECHANISMS IN CMOS DOWNCONVERTERS
The source voltage is related to the input voltage by the following Volterra series expansion:
H
H
(A.2)
405
sinusoid
. The RF DSB-SC input current signal
.
is
The – characteristic of device M is expressed by means
of its Taylor series expansion
is the linear transfer function. H
is found
where H
by applying the Kirchoff law at the source node, i.e.,
Y
(A.3)
and
, as defined in (9),
By means of (A.1)–(A.3),
can be derived. After some simplifications
(B.1)
The solution is found by applying the Kirchoff law
(with
) and expressing
versus the
applied input signals by means of a Volterra series expansion
H
Y
(B.2)
Y
(A.4)
(A.5)
Y
Y
.
with
Assuming no mismatches, the small-signal coefficients are
and
) and
is
equal (
null. Low-frequency output intermodulation products are found
. As a consequence, in the
for
. Moreover, if the LC
LC-degenerated topology, Y
,
tank resonates approximately at the input frequencies
Y
. Under these assumptions
Y
, the third-order nonlinear transfer function
with linear input voltage and input current square, is only
. In particular,
responsible for sidebands around
determines two sidebands in the source
, all
voltage spectrum. To determine
the transfer functions in (B.2) need to be determined. After
calculations and simplifications
(B.3)
pseudodifferential and LC degenerated
Y
fully differential.
(A.6)
is null, its sensitivity to mismatches is not.
Although
,
In presence of mismatches,
,
, and
,
variable is either
or . Taking the
where the generic
derivative of (A.5) with respect to gives
pseudodifferential
LC degenerated
fully differential
Let us first consider the sidebands in the source voltage
around odd harmonics of the LO frequency, due to the
square-wave offset signal. Expanding the square-wave offset
in a Fourier series and by means of (B.3), the sidebands in
the source voltage spectrum are found. The corresponding
sidebands in the capacitor current spectrum are given by
(B.4)
.
where
The current is downconverted at baseband by the switching
stage with a transfer function well approximated by a square
wave toggling between 1 and 1, in phase with the gate
voltage. The upper and lower sidebands add to each other after
downconversion. The resulting output intermodulation current
is given by
(A.7)
Assuming the devices working in strong inversion, Table I
follows.
APPENDIX B
Referring to the circuit in Fig. 12, we look for second-order
, due
intermodulation components in the output current
, with
. We conto sidebands generated around
: a square wave, toggling
sider two applied input signals
with period
, and a large rectified
between 0 and
Cv
(B.5)
From (B.3) and (B.5), assuming
, (22) follows.
The sidebands around even harmonics of the LO frequency
are due to the equivalent rectified sinusoid at the gate. The
Fourier series expansion of the gate voltage is reported in (16).
406
IEEE JOURNAL OF SOLID-STATE CIRCUITS, VOL. 38, NO. 3, MARCH 2003
(B.6)
Again, from (B.3), the source voltage is found. The current
induced in the tail capacitor is given by (B.6), shown at the top
of the page.
Multiplying (B.6) times (19) gives the output intermodulation
current
[9] D. Manstretta, R. Castello, F. Gatta, P. Rossi, and F. Svelto, “A 0.18-m
CMOS direct conversion receiver front-end for UMTS,” in IEEE Int.
Solid State Circuits Conf. Dig. Tech. Papers, Feb. 2002, pp. 240–241.
[10] D. Manstretta, R. Castello, and F. Svelto, “Low 1/f noise CMOS active
mixers for direct conversion,” IEEE Trans. Circuits Syst. II, vol. 48, pp.
846–850, Sept. 2001.
(B.7)
Danilo Manstretta (M’98) was born in Broni, Italy,
in 1973. He received the Laurea degree (summa cum
laude) and the Ph.D. degree in electrical engineering
and computer science from the University of Pavia,
Pavia, Italy, in 1998 and 2002, respectively. During
his studies, he worked on CMOS RF front-end circuits for wireless applications.
In 2001, he joined Agere Systems, Berkeley
Heights, NJ, as a Member of Technical Staff. His
current research interests are in the field of highly
integrated transceivers for WLANs.
From (B.3) and (B.7), assuming
and
, (23) follows.
ACKNOWLEDGMENT
The authors wish to thank M. Paparo and M. Bonaventura
of STMicroelectronics for technology access, and R. Castello,
F. Gatta, and P. Rossi, who worked on the UMTS receiver
realization.
REFERENCES
[1] A. A. Abidi, “Direct conversion radio transceivers for digital communications,” IEEE J. Solid-State Circuits, vol. 30, pp. 1399–1410, Dec.
1995.
[2] B. Razavi, “Design considerations for direct-conversion receivers,”
IEEE Trans. Circuits Syst. II, vol. 44, pp. 428–435, June 1997.
[3] S. Laursen, “Second-order distortion in CMOS direct conversion receivers for GSM,” in Proc. Eur. Solid State Circuits Conf. (ESSCIRC),
Sept. 1999, pp. 342–345.
[4] L. Sheng, J. C. Jensen, and L. E. Larson, “A wide-bandwidth Si/SiGe
HBT direct conversion sub-harmonic/downconverter,” IEEE J. SolidState Circuits, vol. 35, pp. 1329–1337, Sept. 2000.
[5] K. Kivekas, A. Parssinen, and K. A. I. Halonen, “Characterization of
IIP2 and dc offsets in transconductance mixers,” IEEE Trans. Circuits
Syst. II, vol. 48, pp. 1028–1038, Nov. 2001.
[6] M. T. Terrovitis and R. G. Meyer, “Intermodulation distortion in currentcommutating CMOS mixers,” IEEE J. Solid-State Circuits, vol. 35, pp.
1461–1473, Oct. 2000.
[7] Y. Tsividis, Operation and Modeling of the MOS Transistor, 2nd
ed. New York: McGraw-Hill, 1999.
[8] H. Darabi and A. A. Abidi, “Noise in RF-CMOS mixers: A simple physical model,” IEEE J. Solid-State Circuits, vol. 35, pp. 15–25, Jan. 2000.
Massimo Brandolini (S’03) was born in Broni,
Italy, in 1977. He received the Masters degree
in electronics engineering from the University of
Pavia, Pavia, Italy, in 2002. He is currently working
toward the Ph.D. degree in the area of CMOS
RF transceivers design in the Microelectronics
Laboratory, University of Pavia.
His current research activity is focused on RF
front-end design for multistandard applications.
Francesco Svelto (S’94–M’98) received the Ph.D.
degree in electronics and computer science from
the University of Pavia, Pavia, Italy, in 1995. His
Ph.D. dissertation focused on low-noise design for
instrumentation.
In 1997, he joined the University of Bergamo,
Bergamo, Italy, as an Assistant Professor and, in
2000, he was appointed an Associate Professor at the
University of Pavia. His current research interests are
in the field of CMOS RF design and high-frequency
integrated circuits for telecommunications.
Dr. Svelto has been a Member of the Technical Committee of the IEEE
Custom Integrated Circuits Conference (CICC) since 2000. He was a Member
of the Technical Committee of the European Solid State Circuits Conference
(ESSCIRC) in 2002.